Vous êtes sur la page 1sur 22

Symmetry, Integrability and Geometry: Methods and Applications SIGMA 8 (2012), 019, 22 pages

Tippe Top Equations and Equations


for the Related Mechanical Systems
Nils RUTSTAM

Department of Mathematics, Linkoping University, Linkoping, Sweden


E-mail: ergoroff@hotmail.com
Received October 21, 2011, in final form March 27, 2012; Published online April 05, 2012
http://dx.doi.org/10.3842/SIGMA.2012.019

Abstract. The equations of motion for the rolling and gliding Tippe Top (TT) are noninte-
grable and difficult to analyze. The only existing arguments about TT inversion are based
on analysis of stability of asymptotic solutions and the LaSalle type theorem. They do not
explain the dynamics of inversion. To approach this problem we review and analyze here
the equations of motion for the rolling and gliding TT in three equivalent forms, each one
providing different bits of information about motion of TT. They lead to the main equation
for the TT, which describes well the oscillatory character of motion of the symmetry axis 3
during the inversion. We show also that the equations of motion of TT give rise to equations
of motion for two other simpler mechanical systems: the gliding heavy symmetric top and
the gliding eccentric cylinder. These systems can be of aid in understanding the dynamics
of the inverting TT.
Key words: tippe top; rigid body; nonholonomic mechanics; integrals of motion; stability;
gliding friction
2010 Mathematics Subject Classification: 70F40; 74M10; 70E18; 70E40; 37B25

The Tippe Top (TT) is known for its counterintuitive behaviour; when it is spun, its rotation
axis turns upside down and its center of mass rises. Its behaviour is presently understood
through analysis of stability of its straight and inverted spinning solution. The energy of TT
is a monotonously decreasing function of time and it appears that it is a suitable Lyapunov
function for showing instability of the straight spinning solution and stability of the inverted
spinning solution [1, 7, 10, 11, 19]. The energy is also a suitable LaSalle function to conclude
that for sufficiently large angular momentum L directed close to the vertical z-direction the
inverted spinning solution is asymptotically attracting [1, 19].
Since the 1950s [9] it is also known how TT has to be built for the inversion to take place,
more precisely that if 0 < < 1 denotes the eccentricity of the center of mass then the quotient
of moments of inertia = II13 has to satisfy the inequality 1 < < 1+. It is also understood
that the gliding friction is necessary for converting rotational energy to potential energy.
When it comes to describing the actual dynamics of TT, there is very little known. In several
papers [5, 7, 16, 22] numerical results of how the Euler angles ((t), (t), (t)) change during
inversion are presented, but an understanding of qualitative features of solutions (as well as
details of physical forces and torques acting during the inversion) remains a rather unexplored
field.
There is also an alternative approach [18, 21] through the main equation of the TT that shows
the source of oscillatory behaviour of (t) and of the remaining variables (t), (t). The main
difficulty in making further progress lies in the complexity of the TTs equations of motion.
The TT is described, in the simplest setting, by six nonlinear dynamical equations for six
variables ((t), (t), (t), 3 (t), x (t), y (t)). Analysis is slightly simplified by the fact that these
equations admit one integral of motion, the Jellett integral, and that the energy function is
monotonously decreasing in time. To make further progress in reading off the dynamical content
of the equations it is useful to study equations in all possible forms, to study special solutions and
2 N. Rutstam

certain limiting cases of TT equations. By special cases we mean reduced forms of TT equations
obtained by imposing time-preserved constraints and some limiting forms of TT equations. For
instance in this paper, we show that TT equations reduce, in a suitable limit, to the equations
for a gliding heavy symmetric top. These reduced and limiting equations are usually simpler to
study and the lessons learned from these cases are useful for better understanding of dynamics
of the full TT equations.
In this paper we review the results for the TT, modelled as an axisymmetric sphere, with the
aim of understanding how a rigorous statement about dynamics of TT can be derived from full
equations for TT and the role of extra assumptions made in earlier works.

1 Notation and a model for the TT


Inversion of the TT can be divided into four significant phases. During the first phase the
TT initially spins with the handle pointing vertically up, but then starts to wobble. During
the second phase the wobbling leads to inversion, in which the TT flips so that its handle is
pointing downwards. In the third phase the inversion is completed so the TT is spinning upside
down. In the last phase it is stably spinning on its handle until the energy dissipation, due to
the spinning friction, makes it fall down. Demonstrations shows that the lifespan of the first
three phases are usually short, with the middle inversion phase being the shortest. The fourth
phase lasts significantly longer.
This empirical description of the inversion behaviour of the TT is comprehensible for anyone
who observes the TT closely, but it lacks precision and clarity when it comes to answering the
question of how this inversion occurs, and under which circumstances.
We model the TT as a sphere with radius R and axially symmetric distributed mass m. It
is in instantaneous contact with the supporting plane at the point A. The center of mass CM
is shifted from the geometric center O along the symmetry axis by R, where 0 < < 1 (see
Fig. 1). This model has been used in many works, for example by Hugenholtz [9], Cohen [5]
and Karapetyan [11, 12]. An alternative model of the TT could be to consider the TT as two
spherical segments joined by a rigid rod, as proposed in [23].


2
O
CM R
s
K0 a
A

Figure 1. Diagram of the TT model. Note that a = R3 Rz.

We choose a fixed inertial reference frame K0 = (X, b Yb , Z)


b with X b and Yb parallel to the
supporting plane and with vertical Z. Let K = (x, y, z) be a frame defined through rotation
b
around Zb by an angle , where is the angle between the plane spanned by X b and Zb and
the plane spanned by the points CM , O and A. The third reference frame is K = (1, 2, 3),
with origin at CM , defined through a rotation around y by an angle , where is the angle
between Z b and the symmetry axis. The frame K is not fully fixed in body since it rotates. We
let 3 be parallel to the symmetry axis.
Tippe Top Equations and Equations for the Related Mechanical Systems 3

The Euler angles of the body relative to K0 are (, , ). The reference frame K rotates with
the angular velocity z w.r.t. K0 and the angular velocity of the frame K w.r.t. K0 is

ref = 2 + z = sin 1 + 2 + cos 3.

The total angular velocity of the TT is found by adding the rotation around the symmetry
axis 3:

= ref + 3 = sin 1 + 2 + ( + cos )3,

and we shall refer to the third component of this vector as 3 = + cos .


For an arbitrary vector B in the rotating frame K the time-derivative is
 
d  B
B = B1 1 + B2 2 + B3 3 = + ref B.
dt t K

Here the first term B1 1 + B2 2 + B3 3 is the time-derivative of each component describing the
change of the vector B w.r.t. the frame K and ref B the change of B due to rotation of K
w.r.t. the frame K.
Let a = R(3 z) be the vector from CM to the point of contact A. Newtons equations
for the TT describe the motion of the CM and rotation around the CM :

ms = F mgz, L = a F. (1.1)

The external friction-reaction force F acts at the point of support A. The angular momentum is
L = I, where I is the inertia tensor w.r.t. CM and is the angular velocity. If we denote the
principal moments of inertia by I1 , I2 and I3 and note that I1 = I2 due to the axial symmetry,
the inertia tensor has the form I = I1 (11t + 22t ) + I3 33t (the superscript t meaning here
the transpose). The motion of the symmetry axis 3 is described by the kinematic equation
3 = 3 = I11 L 3.
We shall in our model of the TT assume that the point A is always in contact with the plane,
t
which can be expressed as the contact criterion z (s(t) + a(t)) 0. Since this is an identity with
respect to time t all its time derivatives have to vanish as well; in particular, z (s + a) = 0.
So the velocity vA (t) = s + a, which is the velocity of the point in the TT that is in
instantaneous contact with the plane of support at time t, will also have zero z-component.
For the force acting at the point A we assume F = gn zgn vA . This force consists of a normal
reaction force and of a friction force Ff of viscous type, where gn and are non-negative. Other
frictional forces due to spinning and rolling will be ignored. Since we have one scalar constraint
equation z (s + a) = 0, the component describing the vertical reaction force gn z is dynamically
determined from the second time-derivative of the constraint. The planar component of F has
to be specified independently to make the equations (1.1) fully determined. In our model we
take Ff = gn vA = (L, 3, s, s, t)gn (t)vA as the planar component.
This way of writing the friction force indicates that it acts against the gliding velocity vA and
that the friction coefficient can in principle depend on all dynamical variables and on time t.
We keep here also the factor gn (t) to indicate that the friction is proportional to the value of
the reaction force and that we are interested only in such motions where gn (t) 0.
These assumptions about the external force are common in most of the literature about the
TT. Cohen [5] states that a frictional force opposing the motion of the contact point is the only
external force to act in the supporting plane. He used this assumption in a numerical simulation
that demonstrates that inversion of TT occurs. This showed the effectiveness of the model and
in later works such as [4, 7, 14, 15, 19], the external force defined in this way is used without
comment. Or [16] discusses an inclusion of a nonlinear Coulomb-type friction in the external force
4 N. Rutstam

along with the viscous friction. A Coulomb term would in our notation look like C gn vA /|vA |,
where C is a coefficient. Numerical simulation shows that this Coulomb term can contribute
to inversion, but has weaker effect. Bou-Rabee et al. [1] use this result to argue for only using
an external force with a normal reaction force and a frictional force of viscous type in the model
of the TT, since the nonlinear Coulomb friction only results in algebraic destabilization of the
initially spinning TT, whereas the viscous friction gives exponential destabilization.
The above model of the external force becomes significant if we consider the energy function
for the rolling and gliding TT
1 1
E = ms2 + L + mgs z.
2 2
When the energy is differentiated the equations of motion yield E = F vA and if the reaction
2 . Thus for this model of a rolling and gliding TT,
force is F = gn z gn vA , then E = gn vA
the energy is decreasing monotonically, as expected for a dissipative system.
During the inversion of the TT the CM is lifted up by 2R, which increases the potential
energy by 2mgR. This increase can only happen at the expense of the kinetic energy T =
1 2 1
2 ms + 2 L of the TT. Analysis of the inversion must address how this transfer of energy
occurs in the context of the friction model. The following proposition, due to an argument made
by Del Campo [6], gives an idea of how it works.
Proposition 1. The component of the gliding friction that is perpendicular to the (z, 3)-plane
is the only force enabling inversion of TT.
Proof . Inversion starts with the TT spinning in an almost upright position and ends with the
TT spinning upside down, so the process of inversion requires transfer of energy from the kinetic
term 12 L to the potential term mgs z. The angular momentum at the initial position L0
and the final position L1 are both almost vertical, and the value has been reduced: |L1 | < |L0 |.
We let the reaction force F be split into FR + Ff || + Ff , where FR = gn z is the vertical
reaction force, Ff || is the component of the planar friction force parallel to the (z, 3)-plane,
and Ff is the component of the planar friction force that is perpendicular to this plane. We
thus have
L = a (FR + Ff || + Ff ),
which implies that
 
L z = a (FR + Ff || ) z + (a Ff ) z = (a Ff ) z < 0.
We see that without the friction force Ff there is no reduction of L z and no transfer of energy
to the potential term. The torque a (FR + Ff || ) is parallel to the plane of support and causes
only precession of the vector L. 
Fig. 2 illustrates this. The rotational gliding of the TT at the contact point A creates an
opposing force, which gives rise to an external torque that reduces the z-component of the
angular momentum L and transfers energy from rotational kinetic energy into the potential
energy. As a consequence, the CM of the TT rises.
Indeed, in Fig. 2a we see that L is closely aligned with the z-axis so the angular velocity is
pointing almost upwards and the contact velocity vA is pointing into the plane of the picture. In
Fig. 2b it is clear that a Ff = a (gn vA ) gives a torque that has a negative z-component.
It is z (a (gn vA )) = (gn R sin )vA 2. This component has larger magnitude in Fig. 2b
when the inclination angle is closer to /2 and then decreases when the TT approaches the
inverted position (Fig. 2c).
This may help us to understand the observed phenomenon that the inversion of TT is fast
during the middle phase but is slower in the initial and in the final phase, since the z-component
of the torque responsible for transfer of energy and inversion is larger in the middle phase.
Tippe Top Equations and Equations for the Related Mechanical Systems 5

z
z
3

3
vA a
a
a
A A A
F f = g n v A

a) b) c) 3

Figure 2. The TT at different inclination angles during inversion.

1.1 The vector and the Euler angle forms of equations of TT


We return to the equations of motion for the TT in vector form
1
ms = F mgz, L = a F, 3 = L 3. (1.2)
I1
t
The contact constraint z (s + a) 0 is a scalar equation which further reduces this system.
As previously mentioned, all time derivatives of the contact constraint also have to vanish
identically. The first derivative says that the contact velocity has to be in the plane at all times:

z (s + a) = z vA = 0.

The second derivative gives that the contact acceleration is also restricted to the plane:

d2
z (s + a) = z vA = 0.
dt2
The vertical component of F can be determined from
d
z F = z (ms + mgz) = m ( a) z + mg.
dt
The planar components of F have to be defined to specify the equations of motion, and this
makes it impossible to distinguish between the friction force and the planar components of the
reaction force. This is why we need to specify the planar part of the reaction force in our model
to make equations (1.2) complete.
The specifics of the functions in the gliding friction, (L, s, s, 3, t) and gn (L, s, s, 3, t), other
than that they are greater than zero are not relevant in the asymptotic analysis, but we can get
the value of gn by taking the second derivative of the contact constraint and using the equations
of motion:
 
d 1 R
z L 3 + L 3 ,

0 = z s + ( a) = (gn mg) +
dt m I1
so that
mgI12 + mR(3z L2 Lz L3 )
gn = ,
I12 + mI1 R2 2 (1 32z ) + mI1 R2 (3z 1)vA 3

where L3 = L 3, Lz = L z, 3z = 3 z and L2 = |L|2 .


6 N. Rutstam

With gn defined this way, we see that no further derivatives of the contact constraint are
d
needed since the second derivative z (s(t) + dt ( a)) = 0 becomes an identity with respect
to time after substituting solutions for s(t), (t) and a(t). The contact constraint determines
the vertical component of s: sz = a z = R(3z 1), and its derivative s z = ( a) z.
Thus the original system can be written as
1
mr = gn vA , L = a F, 3 = L 3, (1.3)
I1

where r = s sz z. This system has 10 unknown variables (L, r, r, 3). Further, if we assume
= (L, s, 3) then this system does not depend explicitly on r, so we effectively have a system
of 8 ODEs with 8 unknowns.
A rolling and gliding rigid body with a spherical shape such as the TT admits one integral
of motion, the Jellett integral

= L a = R(Lz L3 ).

It is an integral since
 
d R
= L a (R3 Rz) L = (a F) a L (L 3) = 0.
dt I1

We can rewrite the reduced equations of motion for the TT with the reaction force in explicit
form
d
(I) = a (gn z gn vA ) , mr = gn vA , (1.4)
dt

using the Euler angle notation. The angular velocity has the form = sin 1 + 2 + 3 3.
We write the gliding velocity of the point of support as vA = x cos 1 + y 2 + x sin 3, where
x , y are components in the 2 z and 2 direction (note here that z vA = 0 as expected). The
equations (1.4) have the following form in Euler angles:

I1 sin 2I1 cos + I3 3 = R( cos )gn y , (1.5)


I1 I1 2 sin cos + I3 3 sin = Rgn sin + Rgn x (1 cos ), (1.6)
I3 3 = Rgn y sin , (1.7)
m x y + R((1 cos ) + 2 sin + sin (( cos ) + 3 )) = gn x , (1.8)


m y + x R(sin (( cos ) + 3 ) + cos (2( cos ) + 3 )) = gn y . (1.9)

By solving this linear system for the highest derivative of each variable (, , 3 , x , y ), we
obtain the system

sin  Rgn x
= I1 2 cos I3 3 Rgn + (1 cos ), (1.10)
I1 I1
I3 3 2I1 cos gn y R( cos )
= , (1.11)
I1 sin
gn y R sin
3 = , (1.12)
I3
R sin
3 (I3 (1 cos ) I1 ) + gn R(1 cos ) I1 (2 + 2 sin2 )

x =
I1
gn x
I1 + mR2 (1 cos )2 + y ,

(1.13)
mI1
Tippe Top Equations and Equations for the Related Mechanical Systems 7
gn y
I1 I3 + mR2 I3 ( cos )2 + mR2 I1 sin2

y =
mI1 I3
3 R
+ (I3 ( cos ) + I1 cos ) x , (1.14)
I1
with
mgI1 + mR cos (I1 2 sin2 + I1 2 ) I3 3 sin2

gn = . (1.15)
I1 + mR2 2 sin2 mR2 sin (1 cos )x
d
Above we have equations in an explicit form, which (if we add the equation = dt ()) can
be written as (, , , 3 , x , y )= f (, , , , , , x , y ). We see directly that the right hand
sides of the equations are independent of and , which shows that we only need to consider
equations for , , , 3 , x and y . So effectively the system (1.3) has 6 unknowns. Further,
since we have the Jellett integral = L a = RI1 sin2 R( cos )I3 3 , the number of
unknowns can be reduced to 5 on each surface of constant value of .

1.2 Equations in Euler angles with respect to the rotating reference system K
The equations of motion derived above were formulated and calculated with respect to the
(1, 2, 3)-system. We can arrive at an equivalent set of equations by formulating them in the
(x, y, z)-system. This is the approach used in [15] and [22]. We first use the relations

x = cos 1 + sin 3, y = 2 and z = sin 1 + cos 3

to get the angular velocity

= (3 cos ) sin x + y + ( + cos (3 cos ))z,

and

L = (I3 3 I1 cos ) sin x + I1 y + (I1 + cos (I3 3 I1 cos ))z.

We need also note that the frame K = (x, y, z) rotates with the angular velocity z. Thus
starting from the equations of motion in the form (1.4)

L = a (gn z gn vA ) , mr = gn vA ,

(where r = sx x+sy y) and taking into account that the time-derivative is taken with respect to the
rotating frame K (which means for an arbitrary vector B in this frame that B = B

t K + zB),
and that the gliding velocity in the (x, y, z)-system is vA = x x + y y. We obtain in the Euler
angles the following equations:
d
(sin (I3 3 I1 cos )) = R(1 sin )gn y ,
dt
I1 + sin (I3 3 I1 cos ) = Rgn sin + R(1 cos )gn x ,
d
I1 + (cos (I3 3 I1 cos )) = Rgn y sin ,
dt
msx = msy gn x ,
msy = msx gn y ,

The advantage of rederiving the equations of motion in this frame is that it becomes easier to
use a version of the gyroscopic balance condition. This condition, as formulated by Moffatt and
Shimomura [14], says that for a rapidly precessing axisymmetric rigid body we will have

:= I3 3 I1 cos 0.
8 N. Rutstam

This approximation, while somewhat applicable to such rigid bodies as a spinning egg, fails for
the TT initially as Ueda et al. have pointed out [22]. This is since the TT starts with initial
angle (0) 0 and the spin velocity about the symmetry axis is large, while a spinning egg
(axisymmetric spheroid) is thought to start lying on the side, which corresponds to initial angle
(0) 2 and small . Experiments have suggested that the condition 0 is approximately
satisfied during some phase of the inversion, and if we use the new variable = I3 3 I1 cos
instead of , then the equations of motion take a simpler form:

sin + cos = R(1 sin )gn y ,


I1 + sin = Rgn sin + R(1 cos )gn x ,
I1 + cos sin = Rgn y sin ,
msx = msy gn x ,
msy = msx gn y .

The challenge is to show that 0 is true in a certain phase of the inversion of TT. We may
note that in the special case of I1 = I3 , then = I3 , so then the gyroscopic balance condition
says that the spin about the symmetry axis becomes small during inversion.
In [14] it has been argued that if we assume the condition 0 for a spheroid, then
it is possible to obtain a simple equation for (t) that can be integrated. Experimental re-
sults (e.g. [5, 22]) show that nutates as it increases, which means that does not increase
monotonously. Thus for the TT an interesting problem is if we can define a mean value for the
angle (t), h(t)i, which increases monotonously for the initial conditions such that TT inverts.
An estimate in [22] claims that a suitably defined mean value of (t) increases during some
time-interval early in the inversion phase. This is claimed under such assumptions that the av-
erage of sin may be neglected and the average of R(1 cos )gn y is negative. Numerical
simulations shows that there exist choices for parameters and initial conditions so that these
assumptions lead to behaviour of solutions that is in agreement with the results of simulations.
After this phase the authors assume that the TT is in a phase where is close to zero, so the
equation for (t) shows that (t) increases on average during the rest of inversion.

1.3 The main equation for the TT


An alternative form of the TT equations is inspired by the integrability of an axially symmetric
sphere, rolling (without gliding) in the plane [2, 3, 20]. The point A is always in contact with
the plane, and an algebraic constraint characterizing this motion is that the velocity of the point
of contact vanishes, i.e. vA = s + a = 0. Thus we have s = a, and F can be eliminated
from (1.1):
 
d d
(I) = ma ( a) + gz , 3 = 3.
dt dt

Notice that these equations differ from (1.4) where assumptions about the reaction force F =
gn z gn vA are necessary to make the system defined, while here the whole force F = ms =
d
m dt (a ) is dynamically determined and does not enter into the equations. In terms of the
Euler angles we obtain three equations of motion for , and 3 :

sin  2
= (I1 cos mR2 ( cos )(1 cos ))
I1 + mR2 (( cos )2 + sin2 )
3 (mR2 (1 cos ) + I3 ) 2 mR2 mgR ,

(1.16)
Tippe Top Equations and Equations for the Related Mechanical Systems 9

3 (I32 + mR2 I3 (1 cos )) 2 cos


= , (1.17)
sin (I1 I3 + mR2 I1 sin2 + mR2 I3 ( cos )2 ) sin
mR2 I1 cos + mR2 I3 ( cos )
 
3 = 3 sin . (1.18)
I1 I3 + mR2 I1 sin2 + mR2 I3 ( cos )2
This system admits three integrals of motion; the Jellett integral

= RI1 sin2 RI3 3 ( cos ),

the energy
1
E = mR2 ( cos )2 (2 + 2 sin2 ) + sin2 (2 + 32 + 23 ( cos ))
 
2
1
+ I1 2 sin2 + I1 2 + I3 32 + mgR(1 cos ),

(1.19)
2
and the Routh integral [20], which can be found by integrating (1.18):
1/2 p
D = 3 I1 I3 + mR2 I3 ( cos )2 + mR2 I1 sin2

:= I3 3 d(cos ),

mR2 I1
where d(cos ) = + ( cos )2 + (1 cos2 ), = I3 and = I3 . If we are given the
expressions for , D and E above then the equations

= 0, D = 0, E = 0,

are (obviously) equivalent to the equations of motion (1.16)(1.18).


Since the equations of motion (1.16)(1.18) do not depend on and , they constitute
effectively a system of three equations for four unknowns (, , , 3 ), where the equation for
is of second order. This is a fourth order dynamical system and the three integrals of motion
(, D, E) reduce the system to one equation. This is done by expressing and 3 as functions
of , D and :
p
D d(cos ) + R( cos )D
3 = p , = p .
I3 d(cos ) RI1 sin2 d(cos )

and by eliminating and 3 from the energy (1.19). We obtain a separable differential equation
in :

E = g(cos )2 + V (cos , D, ), (1.20)

where g(cos ) = 21 I3 ((( cos )2 + 1 cos2 ) + ),


 p
1 ( d(z) + R( z)D)2 (( z)2 + )
V (z = cos , D, ) = mgR(1 z) +
2I3 d(z) R2 2 (1 z 2 )

2 2 2 p
+ D ((1 z ) + 1) + D( z)( d(z) + R( z)D) ,
R

with d(z) = + ( z)2 + (1 z 2 ) > 0.


The function V (z, D, ) is the effective potential in the energy expression. This potential is
and convex for in the interval (0, ). We have that V (cos , D, ) as 0,
well defined
d(1)
if D
6= R(1) . Between these values, V has one minimum, and the solution (t) stays between
the two turning angles 0 and 1 determined from the equation E = V (cos , D, ) (i.e. when
= 0).
10 N. Rutstam

The effective potential for a rolling axisymmetric sphere was derived as early as [3]. In
Karapetyan and Kuleshov [12] the same potential is found as a special case of a general method
for calculating the potential of a conservative system with n degrees of freedom and k < n
integrals of motion. The potential for the rolling TT in the above form was derived and analysed
in [8].
The dynamical states of the rolling TT, as described by the triple (, D, E), are bounded be-
low by the minimal value of the energy E, that is the surface given by F (D, ) = min V (z, D, ).
z(1,1)
Varying the values for and D, the minimal surface defined by this equation can be generated.
In Fig. 3 we have taken the values of the parameters m = 0.02, R = 0.02, = 0.3, I1 = 235
625 mR
2

and I3 = 52 mR2 . A similar surface showing the stationary points min (, D) for values of
and D has also been discussed in [12]. Points on and above this surface correspond to solutions
to equation (1.20) for the rolling TT. In particular, points on the surface define precessional
motions, and along the marked lines we have singled out the degenerate cases of precessional
motion, the spinning motion where = 0 and = .

F(D, )

6
5
4
5
3
2
3
1 x 10
0
0 = (final) =0 (initial)
5
0
D
3 5
x 10 5

Figure 3. Diagram for the minimal surface F (D, ).

An inverting TT starts from a state close to a spinning, upright position and ends at a state
close to a spinning, inverted position. If we consider a trajectory giving the motion of an
inverting TT, then in Fig. 3 it has to stay in a vertical plane of = const. It starts close above
the line given by = 0 at the minimal surface and ends close above the line given by = .
For better understanding of dynamics of the TT we can study time dependence of functions
that are integrals for when the gliding velocity is zero. For very small vA one may expect that
these functions are changing slowly and are approximate integrals of motion.
We already know that satisfies = 0 for a rolling and gliding TT. For the total energy
of TT, we know that E = F vA 0. However we shall consider the expression for the rolling
energy of the TT

1 1
E(L, 3, s) = m( a)2 + L + mgs z.
2 2
This is not the full energy, but the part not involving the quantity vA . Differentiation yields:
 
d d 1
E(L, 3, s) = E mvA (vA 2( a))
dt dt 2
Tippe Top Equations and Equations for the Related Mechanical Systems 11

= F vA vA (ms) + mvA ( a) = mvA ( a). (1.21)


p
For the rolling and gliding TT the Routh function D(, 3 ) = I3 3 d(cos ), also changes in
time:
d mR sin m
D(t) = p (x + y ) = q (z a) vA . (1.22)
dt d(cos ) d(z 3)

We find from the expressions of and D(t) that


p
D(t) d(cos ) + R( cos )D(t)
3 = p , = p , (1.23)
I3 d(cos ) I1 d(cos ) sin2

and eliminate and 3 in the modified energy (same as equation (1.19)) to get the Main
Equation for the Tippe Top (METT):

E(t) = g(cos )2 + V (cos , D(t), ). (1.24)

It ostensibly has the same form as equation (1.20), but now it depends explicitly on time through
the functions D(t) and E(t). This means that separation of variables is not possible, but it is
a first order time dependent ODE easier to analyze than (1.21) and (1.22), provided that we
have some quantitative knowledge about the functions D(t) and E(t).
For generic values of and D the effective potential V (cos , D(t), ) has a single minimum.
Similarly, as in the pure rolling case, equation (1.24) describes an oscillatory motion of the
angle (t) in the continuously deforming potential. For inverting solutions of TT, = L0 R(1
) = L1 R(1+) (where L0 and L1 are the values p of L for = 0 and = , respectively), so L1 =
L0 1 2 to D = L 1
p
1+ and D changes from the value D 0 = L0 + (1 ) 1 0 1+ + (1 + )2 .

Figure 4. Evolution of V (cos , D(t), ), (0, ), for D between D0 and D1 . The evolution of the
minimum value of V (cos , D(t), ) is also marked.

Fig. 4 shows how V (cos , D(t), ) deforms for (0, ), D (D1 , D0 ). We take here the
same physical parameters m = 0.02, R = 0.02, = 0.3, I1 = 235 2 2 2
625 mR and I3 = 5 mR as has
been used for Fig. 3. We let the value of be 10 times a threshold value (more on this in the
next section). It is apparent here that the minimum value of the potential goes from being close
to = 0 when D is close to D0 and close to = when D is close to D1 .
12 N. Rutstam

The TT satisfying the contact criterion is still a system of 6 equations with 6 unknowns, but
for these equations we have defined three functions of time , D(t) and E(t), and have shown
that the equations of motion (1.5)(1.9) are equivalent to the system
d
(, , , 3 ) = 0,
dt
d m mR sin
D(, 3 ) = q (z a) vA = p (x + y ),
dt d(cos )
d(z 3)
d
E(, , , 3 ) = m( a) vA
dt 
= mR sin (( cos ) + 3 )(x + y ) + (1 cos )(x y ) ,
d
mr = gn vA . (1.25)
dt
If we consider the motion of the TT as being determined by the three functions (, D(t), E(t))
satisfying these equations and connected by the METT, a useful method to investigate the
inversion of TT crystallizes.
This system shows that the functions (D(t), E(t)) allows us to analyze the equations of
motion. If we have D(t) and E(t) given, then integrating the METT gives us (t). With this
information we can find the functions (t) and 3 (t) from (1.23). The equations containing
derivatives of D(t) and E(t) in (1.25) are two linear differential equations for the velocities x (t)
and y (t).
Thus the METT enables us to qualitatively study properties of a class of solutions that
describes inverting solutions of the TT.

2 Special solutions of TT equations


We have already considered one class of special solutions, the pure rolling solutions satisfying
the nongliding condition vA = s + a = 0. They are given by quadratures by solving the
separable equation (1.20).
A special subclass of rolling solutions to TT are solutions of the TT where an explicit assump-
tion about the reaction force is taken (F = gn z gn vA ). This means for rolling TT that we
have now two conditions, vA = 0 and F = gn z. Note that the dynamically determined reaction
d
force for the rolling TT, F = mgz m dt ( a), is not vertical in general, so the condition
F = gn z is a further restriction for rolling solutions of the TT equations.
These solutions play a central role in understanding the inversion of TT because they belong
to the asymptotic LaSalle set {(L, 3, s) : E(L, 3, s) = 0}, which attracts solutions of the TT
equations.
Under the pure rolling condition the external force, as given by our model, is purely vertical:
F = gn z. The system of equations becomes
1
mr = 0, L = Rgn 3 z, 3 = L 3, (2.1)
I1
where the third coordinate of the first equation is determined by msz = gn mg. Further
reduction using the pure rolling condition r + ( a)x,y = 0 restricted to the plane of support
yields the autonomous system
1
L = Rgn 3 z, 3 = L 3, (2.2)
I1
d
along with the constraint dt ( a)x,y = 0 (the subscripts x, y indicate that the vector is
restricted to the supporting plane). But as shown in [7] and [19], we have:
Tippe Top Equations and Equations for the Related Mechanical Systems 13

d
Proposition 2. For the solutions to (2.2) the constraint dt ( a)x,y = 0 can be written as

z (L 3) = 0, a = 0.

These conditions define an invariant manifold of solutions to equations (2.2). Thus for the
solutions of the rolling TT under the assumptions of our model we have that: the vectors L, z
and 3 lie in the same plane and the angular velocity is parallel to the vector a. This in turn
implies that the CM remains stationary (s = 0). These are the situations where the TT is either
spinning in the upright position, spinning in the inverted position or the TT is rolling around
the 3-axis in such a way that the CM is fixed (tumbling solutions).
In terms of the Euler angles, (2.1) gives rise to three equations of motion and two constraints:

I1 sin 2I1 cos + I3 3 = 0,


I1 + I3 3 sin I1 2 cos sin = Rgn sin ,
I3 3 = 0,
mR (1 cos ) + 2 sin + 2 sin ( cos ) + 3 sin = 0,


mR 3 sin + 3 cos + sin ( cos ) + 2 cos ( cos ) = 0, (2.3)

and the quantity gn is determined by (1.15) with x = 0.


After substituting the equations of motion into the constraint equations, we end up with the
following conditions:

sin I1 (2 + 2 sin2 ) + 3 (I1 I3 + I3 cos ) Rgn (1 cos ) = 0,




3 (I3 + (I1 I3 ) cos ) = 0. (2.4)

Using these conditions, we can determine admissible types of solutions to (2.1). As shown in [21],
these constraints imply sin = 0, or that is constant for these solutions to the rolling TT.
Note that this is the condition z (L 3) = 0 from Proposition 2 when expressed in terms of
Euler angles.
We see that for = 0 or = we have the upright or inverted spinning TT, and for constant
(0, ) the first constraint equation (2.4) gives

I1 2 sin2 + 3 (I1 I3 + I3 cos ) Rgn (1 cos ) = 0. (2.5)

The first equation of (2.3) implies that = 0 (i.e. is constant) and the second equation
of (2.3) gives

I3 3 I1 2 cos = Rgn . (2.6)

By eliminating gn between (2.5) and (2.6) we obtain I1 2 ( cos ) + I1 3 = 0. Here 6= 0.


The opposite would lead to gn sin = 0 in (2.3) (thus contradicting the assumption sin 6= 0)
and so

3 = (cos ).

Using again (2.6) with the expression for gn we get a second equation for and 3 :

mgRI1 mR2 2 sin2 (I1 2 cos I3 3 )


I3 3 I1 2 cos =
I1 + mR2 2 sin2
I3 3 I1 2 cos = mRg.

From this relation we easily deduce that gn = mg for rolling solutions to TT with vertical
reaction force (this is obvious if sin = 0). We summarize these results in a proposition.
14 N. Rutstam
d
Proposition 3. The solutions to the system (2.2) under the constraint dt ( a)x,y = 0 are
either spinning solutions = 0, , or tumbling solutions characterized by = 0, (0, ),
where and 3 are determined by the system

3 + ( cos ) = 0, I3 3 I1 2 cos = mRg. (2.7)

Note that if we go back to the angular velocity , the first equation above becomes + = 0.
This is the condition found by Pliskin [17] where it came up by considering a precessing TT
with stationary CM . These are the tumbling solutions.
We can formally solve (2.7) to obtain equations for the angular velocities as functions of the
inclination angle:
s s
mRg mRg
= , 3 = ( cos ) . (2.8)
I3 ( cos ) + I1 cos I3 ( cos ) + I1 cos

These equations together with the value of the Jellett integral = R(I1 sin2 I3 3 (cos )),
determine signs in equations (2.8). The values of the parameters , I1 and I3 give restrictions
whether these equations are defined for all values of cos in the range (1, 1).
The right hand side of equations (2.8) are real for a full range of cos if I3 +(I1 I3 ) cos > 0
for all cos (1, 1). If I1 > I3 we see by setting cos = 1 that I1 < I3 (1 + ). If on the
other hand I1 < I3 then by setting cos = 1 we have the condition I1 > I3 (1 ). So if the
parameters satisfy:

1 < < 1 + , (2.9)

where = I1 /I3 , then and 3 in (2.8) are defined and real for all cos (1, 1). For
outside this interval, and 3 will be real for cos belonging to a subinterval of (1, 1), which is
characterized by the parameters and . Given a value of Jelletts integral , we can investigate
these intervals for when tumbling solutions exist and the number of tumbling trajectories for
that value (see [19]).
In [7], the relative stability (in the sense of Lyapunov) of a given spinning and tumbling
solution is derived as a relation between the integral (specifying initial conditions for the TT
in terms of the angular momentum) and physical characteristics of the TT.
Such a solution is stable if, given a value for , E () has a minimum, where E p () is (1.20)
under the conditions = 0 and 3 + ( cos ) = 0 (which is equivalent to d(cos )(
cos ) + DR( sin2 + ( cos )2 ) = 0):
 p  
E () = V cos , D = d(cos )( cos ) / R sin2 + ( cos )2 ,


2
= + mgR(1 cos ),
2R2 (I1 (1 cos2 ) + I3 ( cos )2 )

which is the form for the potential for steady motions of an axisymmetric sphere, derived
in [10, 11].
mgR3 I3 (1+)2
In particular, if || is above the threshold value
1+
, only the inverted spinning
position cos = 1 is stable [7, 18]. This analysis thus specifies how fast a given TT should be
spun to make solutions to (2.1) stable.
We have characterized the solutions to (1.3) such that vA = 0 and F = gn z. Clearly solutions
to (2.2) are also solutions to the system (1.2). To be precise, they are part of the precessional
solutions, for which the angle is constant. The question is whether these are also asymptotic
solutions to the system (1.3).
Tippe Top Equations and Equations for the Related Mechanical Systems 15

In [19] a theorem of LaSalle type [13] has been formulated to confirm this fact. It is shown
that each solution to (1.3) satisfying the contact criterion z (a + s) = 0 and such that gn (t) 0
for t 0 approaches exactly one solution to (2.1) as t . This is because the trajectories
drawn by the solutions to (2.1) constitute an invariant submanifold to the system (1.3) and
belong to the largest invariant set in the LaSalle set {(L, 3, s) : E(L, 3, s) = 0}. Thus the
solution set to (2.1) can be seen as an asymptotic set for the solutions of the TT equations.
Analysis of special solutions of the TT when vA = 0 thus leads to the most important
conclusion about TT behaviour. The application of LaSalles theorem says that trajectories of
the system (1.3) approach the asymptotic set, which consists of spinning and tumbling solutions.
From the analysis of the equations of motion in the Euler angle form it follows that tumbling
exist for all cos (1, 1) if (2.9) is satisfied. If moreover is above the threshold
solutions
mgR3 I (1+)2
value 3 , only the inverted position is stable. Thus we have conditions for when an
1+
inverted spinning solution is the only attractive asymptotic state in the asymptotic LaSalle set.
Therefore a TT satisfying these condition has to invert. But analysis of dynamics of inversion
is still in its infancy.

3 Reductions of TT equations to equations


for related rigid bodies
In this section we present two reductions of the TT equations that, remarkably, describe motion
of simpler rigid bodies. They are given by differential equations that are simpler but nevertheless
retain certain essential features of the TT equations. They are interesting in their own right
and may also work as a testing ground for ideas of how to analyse equations of TT.

3.1 The gliding HST


The gliding heavy symmetric top (HST) equations represent an axially symmetric top that is
allowed to glide in the supporting plane.
The movements of the gliding HST are described using the same reference systems as before.
In Fig. 5 we have placed K = (x, y, z) with origin at the contact point A, but this is not essential
in our calculations. We let s be the position vector for the CM in the inertial system K0
and a = l3 is the vector from CM to A. We shall assume the contact criterion, i.e. that
t
z (s(t) + a(t)) 0.

CM
z
s l
1
K0 y
vA A
x

Figure 5. Diagram of the gliding HST. Note that x is in the plane spanned by z and 3.
16 N. Rutstam

We consider equations of motion for the gliding HST with respect to the CM , and the only
force creating a torque on CM is the external force F acting at A and having the moment
arm a. For the force applied at A we have F = gn vA + gn z. The equations of motion (1.4)
specialize to:

L = (l3) (gn z gn vA ) , mr = gn vA , 3 = 3, (3.1)

This is a system of 11 equations for the variables (L, r, r, 3) and for the value of the normal
d d2
force gn . The constraint 0 = z (s + a) and its time derivatives dt (s + a) z = 0, dt 2 (s + a) z = 0
determine sz , sz and

mgI1 ml(cos (I1 2 sin2 + I1 2 ) I3 3 sin2 )


gn = .
I1 + ml2 sin2 + ml2 x cos sin

System (3.1) rewritten in Euler angles and solved w.r.t. (, , 3 , x , y ):


1
I1 2 sin cos I3 3 sin + lgn x cos + lgn sin ,

= (3.2)
I1
1 
= I3 3 2I1 cos + lgn y , (3.3)
I1 sin
3 = 0, (3.4)
l sin
I3 3 cos + I1 2 + 2 sin2 lgn cos
 
x =
I1
gn x
I1 + ml2 cos2 + y ,

(3.5)
mI1
lI3 3 I1 + ml2
y = gn y x . (3.6)
I1 mI1

We see that L3 = I3 3 is an integral of motion, due to rotational symmetry about the 3-axis.
The HST with fixed supporting point A also admits

Lz = LA z = (L + ma ( a)) z = I1 sin2 + I3 3 cos

as an integral of motion (where I1 = I1 + ml2 is the moment of inertia w.r.t. A). For the gliding
HST however, this quantity is not an integral of motion since:
  
d d
Lz = LA z = (L + ma ( a)) z = (a F) z + m a ( a) z
dt dt
 
d
= (a (ms + mgz)) z + ma ( a) = m(a vA ) z
dt
I3 sin 2 I sin
= ml 3 + 1 lgn y
I1 I1
(according to (3.5) and (3.6)).
Due to the presence of the frictional force, the gliding HST is not a conservative system. So
the differentiation of the energy function E = 12 ms2 + 12 L + mgs z gives the same result
as for the TT: E = F vA . The derivative of the modified energy function, i.e. the part of the
d
energy not involving vA , satisfies dt E(t) = mvA ( a), as for TT (see (1.21)).
With the properties of the three functions L3 , Lz (t) and E(t) established, we see that the
equations of motion (3.2)(3.6) are equivalent to the system of differential equations
d
L = 0,
dt 3
Tippe Top Equations and Equations for the Related Mechanical Systems 17

d I3 sin 2 I sin
Lz (t) = m(z a) vA = ml 3 + 1 lgn y ,
dt I1 I1
d
E(t) = m( a) vA = ml((x y ) cos + sin (y + x )),
dt
d
mr = gn vA ,
dt
in the same way as the system (1.25) is equivalent to equations (1.5)(1.9). If we substitute the
L L (t)L cos
expressions for 3 = I33 , = z I sin32 , into the expression for the modified energy function
1
we get, analogously to (1.24), the Main Equation for the gliding HST (MEgHST):

I1 2 L23 (Lz (t) L3 cos )2 1


E(t) = + + 2 + mgl cos = I1 2 + V (cos , L3 , Lz ). (3.7)
2 2I3 2I1 sin 2

3.2 Transformation from TT to gliding HST


The gliding HST equations have been introduced here because they appear as a limit of the TT
equations when R = l and R 0. This also transforms integrals of TT into the integrals of
the gliding HST as well as METT (1.24) into MEgHST (3.7).
The vector a connecting the center of mass to the point of contact is R(3 z) for the TT
system, and is l3 for the gliding HST. To describe the transformation from TT to gliding
HST we think of the body of TT being stretched (still maintaining its axial symmetry and
the spherical bottom), and the spherical part being shrunk to a point. So it looks like a ball-
point pen during this transformation. The center of mass thus moves up along the 3-axis and
the radius of the sphere becomes small. We take the limit R 0 subjected to the condition
R = l (see Fig. 6). For the vector equations, the thing that changes is the vector a. We

CM
CM
a
a
CM
a
A A
A

Figure 6. Transformation from TT to gliding HST: R = l, R 0.

directly see that the equations of motion for the TT,

ms = F mgz, L = R(3 z) F, 3 = 3,

become the equations of motion for the gliding HST:

ms = F mgz, L = l3 F, 3 = 3,

since a = R3 Rz = l3 Rz l3 as R 0.
The difference between these dynamical systems is that the TT is allowed to glide and roll
in the plane, whereas the gliding HST glides and rotates. In both cases the body in question
18 N. Rutstam

satisfies the contact constraint. We can also see the effects of the transformation applied to the
equations of motion for TT (1.10)(1.14) expressed in the Euler angles. If we substitute R
with l everywhere in the equations and then let R 0, we arrive at equations (3.2)(3.6).
The limiting form of Jelletts integral and of Rouths function is more tricky and does not
immediately lead to L3 and Lz for the gliding HST. This is not surprising since and D are
derived for a spherical body.
For the Jellett integral we get

= lI3 3 + R(I1 sin2 + I3 3 cos ) = lL3 + O(R) lL3 .

This transformation is also obvious in vector notation, since = L a = R(Lz L3 ) lL3


as R = l, R 0. For the Routh function we expand with respect to small R:
s
I1 ml2 2mlR mR2
 
p I1
D = I3 3 d(cos ) = I3 3 + + cos + sin2 + cos2
I3 I3 I3 I3 I3
mR2  1/2
 
p
2
2ml cos 2 2
= 3 I3 (I1 + ml ) 1 + R + I1 sin + I3 cos
I1 + ml2 I1 + ml2
 
ml cos
= 3 I3 I1 1 + R + O R2 .
p 
I1

Combining this with ,

I1 l I3 I1 D
p 
1
= lI I3 3 + RI1 (I1 sin2 + I3 3 cos )
I1 R I1 R 1
 
ml cos 2

l3 I3 I1 1 + R +O R
I1
 
1 2 2
 2

= R(I1 I1 sin + I1 ml I3 3 cos ) + O R
I1 R

= I1 sin2 + I3 3 cos + O(R) = Lz + O(R),




we see that it tends to Lz as R 0. We have thus retrieved from and D both functions L3
and Lz .
Slightly more elaborate is the calculation that the METT (1.24) is transformed to
MEgHST (3.7). It requires careful expansion of the effective potential V (cos , D(t), ) with
respect to the small variable R (see [21]).

Proposition 4. The METT equation E(t) = g(cos )2 + V (cos , D(t), ) becomes, under the
transformation R = l, R 0, the MEgHST:

I1 2 L2 (Lz (t) L3 cos )2


E(t) = + 3 + + mgl cos .
2 2I3 2I1 sin2

The MEgHST has simpler effective potential

L23 (Lz (t) L3 cos )2


V (cos , Lz (t), L3 ) = + + mgl cos
2I3 2I1 sin2

than METT, but retains the same structure.


Tippe Top Equations and Equations for the Related Mechanical Systems 19

3.3 The gliding eccentric cylinder


It is also possible to reduce the TT equations so they describe an eccentric rolling and gliding
cylinder (glC). These equations are a special case of the TT equations when = D = 0.
We think of an eccentric cylinder, homogeneous in the y-direction, rolling and gliding on
a supporting plane in the direction of the x-axis (the cylinder is considered to be long). We
could also see it as a model for an eccentric wheel rolling along the x-axis in the (x, y)-plane. It
has mass m and radius R. The center of mass CM is shifted with respect to the symmetry axis
of the cylinder by R, 0 < < 1. We let the system K = (1, 2, 3) have its origin in CM , where
the 3-axis points along the line determined by CM and the center of the cross-section circle in
the direction from CM toward this center. 1 is orthogonal to 3 and 2 = 3 1 is parallel to the
y-axis (see Fig. 7).


2
O
CM R
s
K0 a
A vA

Figure 7. Diagram of an eccentric cylinder.

We let be the angle between the z-axis and the 3-axis. This will be the single Euler angle
since the cylinder is only rolling in one direction, so that = 2. Further, we let A denote the
point of contact of the cylinder with the supporting line, so that the vector from CM to A is
a = R(3 z) as for the TT.
As we have done previously, we consider the external force acting on the cylinder to be
F = gn z gn vA , where vA = x x is the velocity of the point A. The equations of motion for
the glC have the familiar form: ms = F mgz and L = a F.
Observe that there is no third equation (for 3) since the axis 2 has constant direction parallel
t
to y. As the contact criterion z (s + a) 0 determines the vertical component of s, we reduce
these equations to

mr = gn vA , L = a F, (3.8)

where r = s sz z is the planar component of the position vector for the CM .


This is a system of only two equations: one for and one for x . We thus proceed directly
to the coordinate form of the equations of motion. Note that L = I = I 2, where I = I2 is
the moment of inertia with respect to the axis 2 going through CM . We get

I 2 = Rgn x (1 cos )2 Rgn sin 2,


mx x + mR(1 cos )x + mR2 sin x = gn x x. (3.9)

The equations of motion in the form solved w.r.t. highest derivative are then
Rgn x Rgn
= (1 cos ) sin ,
I I
R2 gn gn x
I + mR2 (1 cos )2 R2 sin .

x = sin (1 cos )
I mI
20 N. Rutstam

The function gn is determined in the same way as (1.15):

mgI + mRI 2 cos


gn = .
I + mR2 2 sin2 mR2 sin (1 cos )x

We directly see that we can get these equations from the equations of motion for TT (1.10)
(1.14) in the special case = 3 = y = 0.This corresponds to = 0, D = 0 for this system,
and the constraint y s = 0 is satisfied (since the cylinder does not move in the y-direction).
These constraints are consistent with the equations (1.10)(1.14).
When we consider the gliding eccentric cylinder, we have three natural special cases that can
be integrated by quadratures:

a) the noneccentric case = 0, vA 6= 0;


b) the nongliding case vA = 0, 6= 0;
c) the nongliding noneccentric case = 0 and vA = 0.

For these cases we examine the limits vA 0 and/or 0 in this model and notice that the
limit vA 0 leads only to the static solution (t) = const.
We consider first the case of the axially symmetric cylinder, = 0. The center of mass and
the geometric center of the cylinder coincides and the vector from CM to A reduces to a = Rz.
We use this in (3.8), so the equations of motion become I = Rgn x , mx + mR = gn x ,
2
and in the solved form: = RI gn x , x = I+mR
mI gn x . These two equations imply that
mRx + (I + mR2 ) = 0, i.e. mRx + (I + mR2 ) is constant. We can also obtain these
equations by letting 0 in (3.9). If we assume that is constant (note here that gn = mg),
the system can be solved explicitly for x (t) and (t):
I+mR2
x (t) = x (0)e I gt ,
mR I+mR
2
gt 1 2
 
(t) = x (0)e I + mRx (0) + I + mR (0) .
I + mR2 I + mR2

If we now let vA 0 in these equations we get the single equation = 0, saying that the
cylinder is turning with constant speed. This describes the motion of a non-eccentric cylinder
rolling without friction.
In the second case we consider equations for a rolling eccentric cylinder. This means we let
vA = 0 in the equations of motion:

ms = F mgz, L = a F.

We eliminate F in the second equation to get


 
d d
(I) = a (ms + mgz) = a m ( a) + mgz .
dt dt

So for the Euler angle we get the nonlinear equation:

mgR sin mR2 2 sin


= ,
I + mR2 (( cos )2 + sin2 )
but this equation is also found through the energy formula, since

0 = E = ms s + L + mg s z
= I + mR2 (sin2 + ( cos )2 ) + 3 mR2 sin + mgR sin .

Tippe Top Equations and Equations for the Related Mechanical Systems 21

So it admits energy as an integral of motion


1
E = 2 I + mR2 sin2 + ( cos )2 + mgR(1 cos ).

2
This is a first order autonomous ODE that can be separated and has a solution expressed by
quadratures. The equation for reduces to the special case of the rolling noneccentric cylinder
= 0 if we let 0 in this equation.
When we take the limit vA 0 in equations (3.9) we get an additional constraint that
F = gn z is vertical so the result has to differ from the previous. The system becomes one
equation of motion with one constraint:

I = Rgn sin , (3.10)


2
mR(1 cos ) + mR sin = 0.
d
The constraint can be rewritten as dt ((1 cos )) = 0, i.e. (1 cos ) is equal to a con-
stant C.
2 cos
The value of gn is deduced from the contact criterion as before: gn = mgI+mRI
I+mR sin2
2 2 . Using
the constraint = 1Ccos and the expression for gn in equation (3.10) we get a polynomial
equation in cos and sin :

sin IC 2 + mR2 C 2 ( cos ) mgR(1 cos )3 = 0.




Thus with the external force F = gn z gn vA we get only solutions with constant when we
let x 0 in the equations of motion (3.9).

4 Conclusions
In this paper we have discussed several different forms of the TT equations: the vector form, the
Euler angle form, the form leading to METT and the form suitable for discussing the gyroscopic
balance condition. Each form gives different insight into the character of motion of TT and
helps to unveil complicated features of TT dynamics.
The vector form helps to explain in simple terms why the gliding friction is necessary for
inversion of TT. It also is the most suitable for asymptotic analysis of TT equations and leads
to formulation of conditions for when the inverted spinning state is the only admissible stable
solution. The conditions for the physical parameters of TT and for the initial conditions seem
to agree remarkably well with the experimentally observed features of TT motion.
Analysis of the Euler angle form of TT equations confirms the picture obtained from the vec-
tor equations. It also shows that the functions (, D, E), being integrals of motion of the rolling
TT, play a special role in understanding the motion of rolling and gliding TT. By eliminating
and 3 from the energy E we obtain the METT equation having similar form as the separation
equation for the rolling TT. From this we can see that during inversion the symmetry axis is
performing fast nutational motion within a nutational band that is moving from the north pole
to the south pole of the unit sphere S 2 .
Since the Euler equations of TT constitute a nonlinear dynamical system of 6 degrees of
freedom we have studied ways of simplifying these equations by taking dynamically invariant
constraints and certain limits of the physical parameters. Remarkably we have found two such
simplifications that even have mechanical interpretation. The limit R = l, R 0 leads to
the gliding HST equations, which is an interesting system in itself, but the equations are also
interesting in the context of the TT equations because they retain all main features of TT but
are simpler for analytical treatment. We have shown also the constraints s y = 0, = 0 and
D = 0 reduces the TT equations to equations for a rolling and gliding eccentric cylinder.
22 N. Rutstam

References
[1] Bou-Rabee N.M., Marsden J.E., Romero L.A., Tippe top inversion as a dissipation-induced instability,
SIAM J. Appl. Dyn. Syst. 3 (2004), 352377.
[2] Chaplygin S.A., On a balls rolling on a horizontal plane, Regul. Chaotic Dyn. 7 (2002), 131148.
[3] Chaplygin S.A., On a motion of a heavy body of revolution on a horizontal plane, Regul. Chaotic Dyn. 7
(2002), 119130.
[4] Ciocci M.C., Langerock B., Dynamics of the tippe top via Routhian reduction, Regul. Chaotic Dyn. 12
(2007), 602614, arXiv:0704.1221.
[5] Cohen R.J., The tippe top revisited, Amer. J. Phys. 45 (1977), 1217.
[6] Del Campo A.R., Tippe top (topsy-turnee top) continued, Amer. J. Phys. 23 (1955), 544545.
[7] Ebenfeld S., Scheck F., A new analysis of the tippe top: asymptotic states and Liapunov stability, Ann.
Physics 243 (1995), 195217, chao-dyn/9501008.
[8] Glad S.T., Petersson D., Rauch-Wojciechowski S., Phase space of rolling solutions of the tippe top, SIGMA
3 (2007), 041, 14 pages, nlin.SI/0703016.
[9] Hugenholtz N.M., On tops rising by friction, Physica 18 (1952), 515527.
[10] Karapetyan A.V., Global qualitative analysis of tippe top dynamics, Mech. Sol. 43 (1995), 342348.
[11] Karapetyan A.V., Qualitative investigation of the dynamics of a top on a plane with friction, J. Appl. Math.
Mech. 55 (1991), 563565.
[12] Karapetyan A.V., Kuleshov A.S., Steady motions of nonholonomic systems, Regul. Chaotic Dyn. 7 (2002),
81117.
[13] LaSalle J.P., Some extensions of Liapunovs second method, IRE Trans. 7 (1960), 520527.
[14] Moffatt H.K., Shimomura Y., Classical dynamics: spinning eggs a paradox resolved, Nature 416 (2002),
385386.
[15] Moffatt H.K., Shimomura Y., Branicki M., Dynamics of an axisymmetric body spinning on a horizontal
surface. I. Stability and the gyroscopic approximation, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci.
460 (2004), 36433672.
[16] Or A.C., The dynamics of a tippe top, SIAM J. Appl. Math. 54 (1994), 597609.
[17] Pliskin W.A., The tippe top (topsy-turvy top), Amer. J. Phys. 22 (1954), 2832.
[18] Rauch-Wojciechowski S., What does it mean to explain the rising of the tippe top?, Regul. Chaotic Dyn. 13
(2008), 316331.
[19] Rauch-Wojciechowski S., Skoldstam M., Glad T., Mathematical analysis of the tippe top, Regul. Chaotic
Dyn. 10 (2005), 333362.
[20] Routh E.J., The advanced part of a treatise on the dynamics of a system of rigid bodies. Being part II of
a treatise on the whole subject, 6th ed., Dover Publications Inc., New York, 1955.
[21] Rutstam N., Study of equations for tippe top and related rigid bodies, Linkoping Studies in Science and
Technology, Theses No. 1106, Matematiska Institutionen, Linkopings Universitet, 2010, available at http:
//urn.kb.se/resolve?urn=urn:nbn:se:liu:diva-60835.
[22] Ueda T., Sasaki K., Watanabe S., Motion of the tippe top: gyroscopic balance condition and stability,
SIAM J. Appl. Dyn. Syst. 4 (2005), 11591194, physics/0507198.
[23] Zobova A.A., Karapetyan A.V., Analysis of the steady motions of the tippe top, J. Appl. Math. Mech. 73
(2009), 623630.

Vous aimerez peut-être aussi